Fatty Acid-Based Biodegradable Polymers - American Chemical Society

Introduction. Fatty acids have been used previously in the development of polymers for ... using two fatty acids: lithocholic acid and ricinoleic acid...
1 downloads 0 Views 932KB Size
Chapter 6

Fatty Acid-Based Biodegradable Polymers: Synthesis and Applications 1

1

1

Downloaded by STANFORD UNIV GREEN LIBR on July 30, 2012 | http://pubs.acs.org Publication Date: January 1, 2009 | doi: 10.1021/bk-2009-1004.ch006

1

Marina Sokolsky-Papkov , Ariella Shikanov , Aviva Ezra , Boris Vaisman , and Abraham J. Domb 1,2

1

2

Department of Medicinal Chemistry and Natural Products, School of Pharmacy - Faculty of Medicine, The Hebrew University of Jerusalem, 91120 Jerusalem, Israel Leonyl Jacobson Chair of Medicinal Chemistry, Affiliated with the David R. Bloom Center for Pharmacy and the Alex Grass Center for Synthesis and Drug Design at The Hebrew University of Jerusalem, 91120 Jerusalem, Israel

Fatty acids have been used previously in the development o f polymers for biomedical applications as they are considered to be inert, inexpensive and biocompatible, as w e l l as possessing low toxicity. This review focuses on use o f different fatty acids for synthesis o f injectable polymers for intended use as drug carriers.

Introduction Fatty acids have been used previously in the development o f polymers for biomedical applications as they are considered to be inert, inexpensive and biocompatible. The main fatty acids w h i c h are used as a base for synthesis o f biomedical polymers (polyanhydrides) are stearic acid (/), erucic acid ( C 2 2 unsaturated fatty acid) dimer (2), bile acid dimer (J), ricinoleic acid (4) and other fatty acids (5), middle long carbon chain ( C 1 2 - 15) dibasic acids, such as dodecanedioic, brassylic acid, tetradecandioic acid and pentadecandioic acid (7). Fatty acids are used in preparation o f microspheres for controlled drug delivery. For example, poly(vinyl alcohol), substituted with lauric, myristic, palmitic, and stearic acids at different substitution degrees was employed for the preparation biodegradable microspheres containing progesterone or indo-

60

© 2009 A m e r i c a n C h e m i c a l Society

In Polymer Degradation and Performance; Celina, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2009.

61

Downloaded by STANFORD UNIV GREEN LIBR on July 30, 2012 | http://pubs.acs.org Publication Date: January 1, 2009 | doi: 10.1021/bk-2009-1004.ch006

methacin (6). Microspheres were prepared by starting from an oil-in-water dispersion containing the polymer and drug in the inner phase, following by solvent extraction method. Microspheres were obtained with high loading efficiency, whose release properties were dependent on the nature o f the acyl substituent and the substitution degree. Fatty acids were incorporated into polymers to influence their physicochemical properties. For example, polyanhydrides based on short chain hydrophobic acids such as sebacic/adipic acids are brittle and hydrolyze under c o m m o n storage conditions. Introduction o f hydrophobic moieties as part o f the polymer chain or as hydrophobic terminals resulted in better hydrolytic stability, pliability and affects biomedical applications such as release profile from the polymers (7). The same approach was used to influence the properties o f w i d e l y used polyester - poly(lactic acid). Polyesters are useful materials for controlled drug delivery since they hydrolyze to hydroxyl acid monomers when placed in aqueous media. Polyesters for drug delivery systems are usually synthesized from lactic/glycolic acids. H o w e v e r enantiomerically pure P L A is a semicrystalline polymer with T g o f 5 5 ° C and T m o f 180°C. The degree o f crystallinity and melting point can be reduced by random incorporation o f other monomers into the P L A chain leading to decrease in crystallization ability (#10) Fatty acids are suitable candidates for incorporation into P L A , as they are natural components and they are hydrophobic and thus may retain an encapsulated drug for longer time period when used as drug carrier (4, 11). This article focuses on the use o f fatty acids in the synthesis and properties o f two types o f the polymers, poly(ester anhydride)s and polyesters.

Poly(ester-anhydrides) Polyanhydrides as a class o f degradable polymers have been used in a number o f applications (12, 13) and in drug delivery. They have been extensively investigated for use in the controlled delivery o f a number o f drugs including chemotherapeutics (14, 15), antibiotics (15), anesthetics (16, 17), and polypeptides (18) in the past 20 years. Recently, polyanhydrides designed for use in the fields o f orthopedics (19) and polymer drugs (20) have been reported. Consequently, curatorial researchers have attached increasingly importance to the polyanhydrides in medical applications. Polyanhydrides prepared from fatty acids are good candidates for the delivery o f hydrophilic drugs due to the desired hydrophobicity o f the natural fatty acids in the main chain o f the polyanhydrides ( / / ) . These polyanhydrides have two series o f acid monomers: one has longer carbon chain, such as dimer erucic or oleic acid; another has shorter carbon chain, such as sebacic acid (2). Fatty acids can be incorporated into the polymer chain by one o f two ways: by

In Polymer Degradation and Performance; Celina, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2009.

62

Downloaded by STANFORD UNIV GREEN LIBR on July 30, 2012 | http://pubs.acs.org Publication Date: January 1, 2009 | doi: 10.1021/bk-2009-1004.ch006

conversion o f monofunctional fatty acids into diacid monomers and including them as repeating units i n the polymer (2) or by using monofunctional fatty acid as chain terminators. The second approach was applied i n the synthesis o f fatty acid terminated polyanhydrides. Polyanhydrides based on sebacic acid, and terminated w i t h oleic, stearic, linoleic or lithocholic acid, or combinations o f several fatty acids were synthesized (21), The general structure o f fatty acid terminated polyanhydrides is shown i n Figure 1.

0

0

n»25-50 Figure

1. Fatty acid terminated polyanhydrides (reproducedfrom R=Oleic, stearic, linoleic, lithocholic acids.

reference 7)

These polyanhydrides all possessed suitable properties including: l o w melting points ( 6 0 - 8 2 ° C ) , biodegradability and pliability. These polyanhydrides were able to retain both hydrophilic ( 5 F U ) and hydrophobic (triamcinolone) drugs. The release o f 5 F U continued for almost 2 weeks and o f triamcinolone for 3 weeks. The incorporation o f fatty acids into a polyanhydride chain was investigated using two fatty acids: lithocholic acid and ricinoleic acid. Lithocholic acid containing polyanhydrides (Figure 2) were prepared by two step synthesis. Polyanhydrides reached molecular weights o f 21000-115000 D a , depending on the polymer composition. Release o f model drugs from these polymers showed sustained release o f 5 F U for almost 3 weeks and triamcinolone for 4 weeks (22). The previously described polyanhydrides still have relatively high melting temperature and the only possible administration route is implantation or injection after heating. The polymer is melted and the drug o f choice is m i x e d into it. The formulation needs to be preheated before administration, w h i c h makes the administration complicated and can cause burns at the injection site. Development o f polyanhydrides w h i c h are pasty or liquid at room temperature is, therefore, desired. R i c i n o l e i c acid is a bifunctional fatty acid containing a hydroxy group along the fatty chain. The presence o f both carboxylic and h y d r o x y l groups allows incorporation o f ricinoleic acid into a polymer backbone by formation o f an ester bond. The synthesis o f poly(ester-anhydride) contains two steps: trans-

In Polymer Degradation and Performance; Celina, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2009.

63

C H

Downloaded by STANFORD UNIV GREEN LIBR on July 30, 2012 | http://pubs.acs.org Publication Date: January 1, 2009 | doi: 10.1021/bk-2009-1004.ch006

Figure 2. Lithocholic

acid containing polyanhydrides reference 21)

3

(reproduced from

esterification reaction between solid poly(sebacic anhydride) and h y d r o x y l group o f ricinoleic acid. The ricinoleic acid hydroxyl group reacts with the anhydride group to form ester bonds and releases a polysebacic ( P S A ) anhydride chain w i t h a carboxylic acid terminus. The next step is repolymerization o f sebacic-ricinoleic acid ( S A - R A ) oligomers using conventional route for polyanhydride synthesis (23). The structure o f poly(sebacic-co-ricinoleic acid) anhydride, p ( S A - R A ) , is shown in Figure 3.

H,C Figure 3. Structure

of poly(sebacic-co-ricinoleic acid) anhydride. from reference 23.

Reproduced

The physicochemical properties o f these polymers are correlated with ricinoleic acid content in the polymer. The higher the ricinoleic content i n the polymer, the lower is the melting temperature o f the polymer. Polymers containing 3 0 % w / w or less o f ricinoleic acid are liquid at body temperature, with melting temperatures o f 3 3 ° C and lower. These polymers are injectable at room temperature through a 2 3 G needle, allowing simple administration o f polymer formulations. The degradability o f the polymers was also evaluated. W h e n incubated in phosphate buffer p H 7.4 the liquid polymer lost 8 0 % o f its weight during 2 weeks o f incubation. L i q u i d polymers solidified in phosphate buffer solution and decomposed completely after 8 weeks in the buffer. D r u g release from both pasty and solid polymers was evaluated using cis-platin as model drug. D r u g release from the pasty polymers was different from drug release from solid formulations. Solid formulations released up to 8 0 % o f the

In Polymer Degradation and Performance; Celina, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2009.

Downloaded by STANFORD UNIV GREEN LIBR on July 30, 2012 | http://pubs.acs.org Publication Date: January 1, 2009 | doi: 10.1021/bk-2009-1004.ch006

64 drug for 3 weeks, while the pasty polymers released the entire drug load in 5 weeks. Solidification can be characterized as an increase in polymer viscosity. The rheological behavior o f the polymer is important to characterize its injectability (24). Poly(sebacic-co-ricinoleic acid) with 7 0 % ricinoleic acid content or higher is liquid at body and room temperature. Poly(sebacic-co-ricinoleic acid) 3:7 (70% ricinoleic acid content) shows the properties o f a non-Newtonian fluid at lower temperatures ( < 3 0 ° C ) , because o f a non-constant shear rate/shear stress relationship, and the polymer can be classified as a pseudoplastic shear-thinning material displaying decreasing viscosity with increasing shear rate. This behavior is important for injectability o f the polymer: as pressure is applied, the polymer paste becomes softer and pumps out through the needle. A t higher temperatures ( > 3 0 ° C ) , poly(sebacic-co-ricinoleic acid) 7:3 acts as a Newtonian fluid and its viscosity is not affected by applied shear rate. After exposure to buffer, the viscosity o f this polymer increased three fold, both at room and body temperature. After exposure to the aqueous medium the polymer shows a pseudoplastic behavior: this may be explained by reorganization o f the polymer chains induced by the exposure to the buffer, which is destroyed at the moment o f turning the spindle. The faster the rotation o f the spindle (higher shear rate), the more the structure is destroyed, and the less the molecules slide together, the lower the viscosity w i l l be. The changes in polymer appearance after exposure to aqueous media were evaluated using light and S E M microscopy. Before exposure to buffer, the polymer is transparent. After exposure to buffer the polymer became opaque, and when it was cut, two different regions were found: the outer region w h i c h is gel and the core, w h i c h appears as a soft solid matrix. S E M analysis o f the interface o f the polymer exposed to buffer showed that a kind o f a rigid network was formed across the polymer sample. This network causes the polymer drop to keep its shape in water. T o evaluate the in-vivo gelation mechanism o f the polymer, mice were injected with three different volumes o f polymer. The polymer remained at the injection site and maintained its shape for 24h post injection as happens when o i l is injected in the subcutaneous space. Based on data presented in this work we can classify poly(sebacic-co-ricinoleic acid) as in-situ organogel forming. The in-situ gelling mechanism o f these pasty polyanhydrides suggests that after injection the polymer remains localized at the site o f administration and releases the loaded drug at the site o f administration. This characteristic o f the polymer formulation allows many limitations o f conventional therapy to be overcome, for example in anti-cancer therapy. When a drug is administered systemically by intravenous, intramuscular, or oral dosing, it is distributed in the body to various organs and tissues perfused with blood, and only a relatively small amount reaches its target tissue. Paclitaxel, a novel antitumor agent, has been shown to be highly cytotoxic and is clinically active against advanced

In Polymer Degradation and Performance; Celina, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2009.

65

Downloaded by STANFORD UNIV GREEN LIBR on July 30, 2012 | http://pubs.acs.org Publication Date: January 1, 2009 | doi: 10.1021/bk-2009-1004.ch006

ovarian and breast cancer. It has a unique mechanism o f action, inducing the assembly o f stable microtubules and inhibiting the depolymerization process. The current clinical dosage form o f paclitaxel consists o f a 1:1 (v/v) mixture o f ethanol and Cremophor E L . This pharmaceutical formulation, however, is associated with a number o f concerns including stability, filtering requirements, and use o f non-plasticized solution containers. Moreover, some o f the side effects, such as severe hypersensitivity reactions, observed after paclitaxel administration are considered to be formulation related (25). Incorporation o f paclitaxel into a degradable polymer matrix allows local treatment o f the tumor, providing therapeutic concentration o f the drug at the tumor area, before or after surgical removal o f the tumor. Implanting a biodegradable device loaded with antineoplastic agent in the cavity created by the tumor provides high local concentration o f the drug k i l l i n g the malignant cells that survived the surgery and also prevents the systemic side effects o f the chemotherapy normally associated with the intravenous administration. Poly(sebacic-co-ricinoleic acid) is a suitable polymer for implantation into a tumor cavity after removal o f the tumor (solid polymers) or injection into localized tumor (pasty polymers). Drugs (paclitaxel or cis-platin) are incorporated into polymer by a trituration technique, resulting in uniform dispersion o f the drug i n polymer matrix. The incorporation o f the drugs does not affect the polymer properties. The rate o f paclitaxel release from the semisolid polymer differs as a function o f the drug loaded in the polymer. A s the percentage o f paclitaxel is higher in the polymer, the rate o f drug release is slower. The formulation containing 5% paclitaxel released 3 3 % o f the incorporated drug in 3 months. The release o f paclitaxel from the solid formulation is faster than from the semisolid. D u r i n g the first 50 days, formulations with 5 and 10% w / w paclitaxel released about 2 0 % o f the loaded paclitaxel, and the formulation containing 2 0 % w / w paclitaxel released only 15% o f the drug content during this time period (6). Cis-platin was incorporated into the p o l y ( S A : R A ) 3:7 and 2:8 ( 5 % w/w). 3 0 % o f the incorporated drug was released in first two days and 8 5 % in the following ten days from both polymers. N o burst release was observed, only 5-7% o f the drug was released in the first day, suggesting that no toxic effects are expected in-vivo (26). These promising results suggested that farther evaluation o f paclitaxel polymer is desirable. The polymer formulations containing anticancer agents (paclitaxel and cisplatin) were evaluated in-vivo in heterotrophic (mouse bladder tumor) and orphotrophic (rat prostate cancer) models. Single administration o f polymerpactlitaxel formulation intratumorally in a mouse bladder tumor model increased the survival rate o f the animals compared to untreated animals and to animals treated w i t h paclitaxel dispersion (conventional administration method) (27). The optimal load o f paclitaxel in the polymer was established as 10% w / w . M i c e treated with this formulation showed median survival rate ( M S R ) o f

In Polymer Degradation and Performance; Celina, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2009.

66 35days (animals were sacrificed at the end point o f the experiment) compared to 16 days for the untreated group and 18 days for animals treated with paclitaxel suspension. 4 0 % o f the mice treated with 10% paclitaxel formulation survived for 77 days. The control groups survived up to 23 days. Treatment with polymer-paclitaxel formulation reduced the tumor size from 3.6g (untreated animals) and 2.5g (paclitaxel suspension) to 0.3 g (10% paclitaxel formulation) (27). In a second study tumor bearing mice were treated with cis-platin formulation (50 u l , 1% cis-platin load). In this study the control animals were sacrificed 18 days post tumor cells inoculation when the tumor reached 3.5 c m . In the group treated with polymer-cis-platin formulation (single injection, ten days post cell inoculation) in 8 out o f 10 mice, the tumor absolutely disappeared during the first 10 days post treatment and d i d not appear until the end o f study (42 days). In the two remaining mice a small nodule reappeared twenty days post treatment, however no increasing o f it growth was seen until the end o f the study. The treatment with the polymer formulation o f paclitaxel and cis-platin had a positive outcome based on both qualitative observations and quantitative measurements o f different parameters. The formulation increased the survival o f mice bearing bladder tumors and decreased the growth rate o f these tumors. The polymer-paclitaxel formulation was also evaluated for treatment o f orthotopic prostate cancer (28). Treatment with the polymer formulation o f paclitaxel (single injection o f 200 u l polymer formulation with 10% w / w load) increased the survival rate o f the rats. Rats treated with parental formulation o f paclitaxel died 25 days post tumor cells inoculation. O n l y one rat in the polymer-paclitaxel group died three weeks post tumor cell inoculation, while all the remaining rats survived until the end point o f the experiment (35 days). The control animals also developed lymph node metastases. N o metastases were found in polymer-paclitaxel treated rats. The treatment with polymer-paclitaxel formulation reduced the prostate volume o f the rats from 14.8 c m (untreated animals) to 0.862 c m , while the volume o f healthy prostate gland injected with 200 u l o f polymer is about 0.4 c m . The polymeric formulation released paclitaxel into local tumor tissues and induced necrosis and reduction o f the tumor mass, while prolonging lifespan in an orthotopic prostate cancer rat model.

Downloaded by STANFORD UNIV GREEN LIBR on July 30, 2012 | http://pubs.acs.org Publication Date: January 1, 2009 | doi: 10.1021/bk-2009-1004.ch006

3

3

3

3

Polyesters Biodegradable polyesters are useful materials for controlled drug delivery (3). They degrade to hydroxy-acid monomers when placed in aqueous medium (7, 29, 30). Polyesters are less sensitive to hydrolysis and their degradation period is longer compared to polyanhydrides. The ester bond is less sensitive to hydrolysis, making preparation, handling and storage o f this type o f polymer easier. Polyesters can be synthesized by polycondensation o f hydroxy acids or

In Polymer Degradation and Performance; Celina, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2009.

Downloaded by STANFORD UNIV GREEN LIBR on July 30, 2012 | http://pubs.acs.org Publication Date: January 1, 2009 | doi: 10.1021/bk-2009-1004.ch006

67 by ring-opening polymerization o f c y c l i c esters (lactones). A wide range o f monomers has been used to produce biodegradable polyesters. The most useful monomers used for polycondensation are lactic, g l y c o l i c , hydroxybutyric, and hydroxycaproic acids. Polyesters o f g l y c o l i c and lactic acids are the main group o f interest due to their long history o f safety (31-33). Enantiomerically pure P L A is a semicrystalline polymer w i t h T g o f about 5 5 ° C and T m o f about 180°C. The degree o f crystallinity and melting temperature o f P L A polymers can be reduced by random copolymerization w i t h other comonomers, leading to the incorporation o f units disturbing the crystallization ability o f the P L A segments. Fatty acids are good candidates for incorporation into the poly(lactic acid) backbone. R i c i n o l e i c acid, as previously described, posses both carboxylic acid and h y d r o x y l functional groups, w h i c h allows its incorporation into polyester chains resulting i n soft hydrophobic polymers. R i c i n o l e i c acid hydrocarbon chain, cis-configuration and side chain may result i n steric hindrance o f the polymer to y i e l d soft or even l i q u i d polyesters. Several polymerization methods have been used for synthesis o f these polymers: ring opening polymerization o f lactide and ricinoleic acid lactones, polycondensation o f lactic and ricinoleic acid and insertion o f ricinoleic acid into poly(lactic acid) by transesterification followed by repolymerization. The polyester structure is shown i n Figure 4.

Figure 4. Structure

of poly(lactic-co-ricinoleic

acid) ester.

L i q u i d polyesters are formed when the feed ratio o f ricinoleic acid is higher than 3 0 % for polycondensation and 5 0 % for transesterification. Polyesters synthesized by ring opening polymerization are solid at r o o m temperature. These polyesters were evaluated as drug carriers for two drugs: 5 F U and triamcinolone i n vitro. B o t h drugs were released for over two weeks (34). A d d i t i o n a l polyesters were synthesized by the polycondensation o f lactic acid and castor o i l (Figure 5) and examined as injectable controlled delivery carriers for cytotoxic drugs (paclitaxel, methotrexate, 5 F U and cis-platin). Polyesters were synthesized using enantiomerically pure lactic acid ( L or D ) or racemic D L lactic acid. Polyesters containing 4 0 % castor o i l are viscous liquids at room temperature and released the incorporated drugs for over two months.

In Polymer Degradation and Performance; Celina, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2009.

68

Downloaded by STANFORD UNIV GREEN LIBR on July 30, 2012 | http://pubs.acs.org Publication Date: January 1, 2009 | doi: 10.1021/bk-2009-1004.ch006

R=L/D/DL Poly(Lactic acid) (ester bond) Figure 5. Poly (lactic acid-co-castor

oil) polyesters.

Conclusions Fatty acid based biodegradable polymers have many biomedical applications. This short review focuses on controlled drug delivery using t w o classes o f the polymers: polyanhydrides and polyesters based on fatty acids as drug carriers. Different polymer types and compositions are summarized showing the potential o f these polymers as drug carriers.

References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13.

G u o , W . X.; Huang, K. X.; Tang, R.; Chi, Q . Polymer 2004, 45, 5743-5748. D o m b , A. J.; Maniar, M. J. Polym. Sci. A. 1993, 31, 1275-1285. D o m b , A. J.; Nudelman, R. J. Polym. Sci. A 1995, 33, 717-725. T e o m i m , D.; Nyska, A.; Domb, A. J. J. Biom. Mat. Res. 1999, 45, 258-267. G o u s i n , S.; Zhu, X.; Lehnert, S. Macromolecules 2000, 30, 5379-5383. Orienti, I.; Zuccari, G.; Luppi, B.; Zecchi, V. J. Microencapsul. 2001, 18, 77-87. T e o m i m , D.; Domb, A. J. J. Polym. Sci. A: 1999, 37, 3337-3344. Grijpma, D. W.; Pennings, A. J. Macromol. Chem. Phys. 1994, 195, 16331647. Grijpma, D. W.; Pennings, A. J. Polym. Bul. 1991, 25, 335-341. Grijpma, D. W.; Nijenhuis, A. J.; Pennings, A. J. Polymer 1990, 31, 22012206. T e o m i m , D.; Domb, A. J. Biomacromolecules 2001, 2, 37-44. U h r i c h , K. E.; Gupta, A.; Thomas, T. T.; Laurencin, C. T.; Langer, R. Macromolecules 1995, 28, 2184-2193. M u g g l i , D. S.; Burkoth, A. K.; Keyser, S. A.; Lee, H. R.; Anseth, K. S. Macromolecules 1998, 31, 4120-4125.

In Polymer Degradation and Performance; Celina, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2009.

Downloaded by STANFORD UNIV GREEN LIBR on July 30, 2012 | http://pubs.acs.org Publication Date: January 1, 2009 | doi: 10.1021/bk-2009-1004.ch006

69 14. Park, E. S.; Maniar, M.; Shah, J. C. J. Control Release 1998, 52, 179-189. 15. Stephens, D.; Li, L.; Robinson, D.; Chen, S.; Chang, H. C.; Liu, R. M.; T i a n , Y. Q.; Ginsburg, E. J.; Gao, X. Y.; Stultz, T. J. Control Release 2000, 63, 305-317. 16. Masters, D. B.; Berde, C. B.; Dutta, S.; Turek, T.; Langer, R. Pharm. Res. 1993, 10, 1527-1532. 17. Shikanov, A.; Domb, A. J.; Weiniger, C. F. J. Control. Release 2007, 117, 97-103. 18. Carino, G. P.; Jacob, J. S.; Mathiowitz, E. J. Control. Release 2000, 65, 261-269. 19. M u g g l i , D. S.; Burkoth, A. K.; Anseth, K. S. J. Biomed. Mater. Res. 1999, 46, 271-278. 20. Erdmann, L.; Macedo, B.; Uhrich, K. E. Biomaterials 2000, 21, 2507-2512. 21. Krasko, M. Y., et al. Polym. Adv. Tech. 2002, 13, 960-968. 22. Krasko, M. Y.; Ezra, A.; Domb, A. J. Polym. Adv. Tech. 2003, 14, 832-838. 23. Krasko, M. Y.; Shikanov, A.; Ezra, A.; Domb, A. J. J. Polym. Sci. A: 2003, 41, 1059-1069. 24. Shikanov, A.; Domb, A. J. Biomacromolecules 2006, 7, 288-296. 25. Shikanov, A.; Vaisman, B.; Krasko, M. Y.; Nyska, A.; Domb, A. J. J. Biom. Mat. Res. A. 2004, 69A, 47-54. 26. Shikanov, A.; D o m b , A. J. Submitted, 2007. 27. Shikanov, A.; D o m b , A. J. Submitted, 2007. 28. Shikanov, A.; D o m b , A. J. Submitted, 2007. 29. Duda, A.; Penczek, S. Polymer 2003, 48, 16-27. 30. Ikada, Y.; Tsuji, H. Macromol. Rap. Commun. 2000, 21, 117-132. 31. K u m a r , N.; Ravikumar, M. N. V.; Domb, A. J. Adv. Drug Deliv. Rev. 2001, 53, 23-44. 32. Penczek, S.; Szymanski, R.; Duda, A.; Baran, J. Macromol. Symp. 2003, 201, 261-269. 33. Seppala, J. V.; Helminen, A. O.; Korhonen, H. Macromol. Biosci. 2004, 4, 208-217. 34. Slivniak, R.; Ezra, A.; Domb, A. J. Pharm. Res. 2006, 23, 1306-1312.

In Polymer Degradation and Performance; Celina, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2009.